ABSTRACT

Let G be a group such that G/Z(G) is finite and simple. The non-commuting, non-generating graphΞ(G) of G has vertex set GZ(G), with edges corresponding to pairs of elements that do not commute and do not generate G. Complementing our previous investigation of this graph for non-simple groups, we show that Ξ(G) is connected with diameter at most 5, with smaller upper bounds for certain families of groups. Using these bounds, we then prove that when G is simple, the diameter of the complement of the generating graph of G has a tight upper bound of 4, with the exception of at most one group with a graph of diameter 5.

1. INTRODUCTION

The generating graph of a group G has been investigated by many authors (for example, [5, 6, 13]) and encodes information about the 2-generation properties of G: its vertex set is G{1} and its edges are the generating pairs for G. The complement of this graph has also been studied (for example, in [38]), and for G non-abelian, its edge set necessarily includes all pairs of commuting elements. The remaining (and more interesting) non-generating pairs are described by the non-commuting, non-generating graphΞ(G) of G, with vertex set GZ(G) and {x,y} an edge if and only if [x,y]1 and x,yG.

Since its introduction in [10], the study of the graph Ξ(G) has been motivated by the importance of the 2-generation properties of groups throughout (abstract and computational) group theory and its applications, and by Cameron’s [9, Section 2.6] hierarchy of graphs. This is a sequence of graphs defined on G (for convenience, we will take the vertex set of each to be G{1}) such that, as long as G is non-abelian, each is a spanning subgraph of the next graph in the sequence. The final three of these graphs are the commuting graph (whose edges are the commuting pairs of elements of G), the non-generating graph (that is, the complement of the generating graph) and the complete graph. The well-studied generating graph is the difference between the final two graphs. It is therefore natural to investigate the next difference, between the non-generating graph and the commuting graph, with all vertices corresponding to central elements deleted (these would otherwise necessarily be isolated). This difference is precisely Ξ(G).

It was proved in [10] that if G is nilpotent (or more generally, if all maximal subgroups of G are normal) and Ξ(G) has an edge, then either the subgraph Ξ+(G) induced by the non-isolated vertices of Ξ(G) is connected with diameter 2 or Ξ(G) itself is connected with diameter 3. In our companion paper [22], we extend these results to the general case where G/Z(G) is not simple. The only additional possibilities here are that G is infinite and Ξ+(G) has diameter 3 or 4, or that Ξ(G) is the union of two connected components of diameter 2 (we provide in that paper a detailed structural description of the finite groups G satisfying this case).

In this paper, we address the case where G/Z(G) is finite and (non-abelian) simple. Our main theorem assumes that G itself is simple. Here, diam(Ξ(G)) denotes the diameter of Ξ(G), which is always at least 2, since no vertex is adjacent to any of its powers. Compared to Theorem 6.1 from the thesis [21] on which this paper is based, our theorem takes into account new information on the maximal subgroups of the monster group M; see Remark 2.7.

 
Theorem 1.1

Let G be a non-abelian finite simple group.

  • Ξ(G) is connected with diameter at most 5.

  • If G=M, then diam(Ξ(G)){4,5}. If instead G is any other sporadic simple group or the Tits group, then diam(Ξ(G))4.

  • For certain simple groups G, Table 1 lists exact values or upper bounds for diam(Ξ(G)).

Table 1.

Exact values or upper bounds for diam(Ξ(G)), for certain simple groups G

GConditionsdiam(Ξ(G))
M11,M12,M22,J22
PSL(2,q)q even or q92
M23,J13
PSL(2,q)q odd and q113
B,PSU(7,2)4
Ann even3
Ann odd4
PSL(n,q),Sz(q)4
G2(q),2G2(q),3D4(q),F4(q),E8(q)q odd4
GConditionsdiam(Ξ(G))
M11,M12,M22,J22
PSL(2,q)q even or q92
M23,J13
PSL(2,q)q odd and q113
B,PSU(7,2)4
Ann even3
Ann odd4
PSL(n,q),Sz(q)4
G2(q),2G2(q),3D4(q),F4(q),E8(q)q odd4
Table 1.

Exact values or upper bounds for diam(Ξ(G)), for certain simple groups G

GConditionsdiam(Ξ(G))
M11,M12,M22,J22
PSL(2,q)q even or q92
M23,J13
PSL(2,q)q odd and q113
B,PSU(7,2)4
Ann even3
Ann odd4
PSL(n,q),Sz(q)4
G2(q),2G2(q),3D4(q),F4(q),E8(q)q odd4
GConditionsdiam(Ξ(G))
M11,M12,M22,J22
PSL(2,q)q even or q92
M23,J13
PSL(2,q)q odd and q113
B,PSU(7,2)4
Ann even3
Ann odd4
PSL(n,q),Sz(q)4
G2(q),2G2(q),3D4(q),F4(q),E8(q)q odd4
 
Remark 1.2

Table 1 shows that there exist non-abelian finite simple groups G with diam(Ξ(G)) equal to 2, 3 or 4. We note that the baby monster group B, the monster group M and PSU(7,2) are the only known simple groups G for which diam(Ξ(G))>3. Indeed, we do not know if infinitely many such groups exist, nor if diam(Ξ(G))=5 is possible.

At the end of each of Sections 3.1, 3.3, 5.1 and 5.2, we provide further examples of diameters achieved by various simple groups, computed using Magma [3] (see Section 2.1 for a discussion of our computational methods). Notably, none of our examples meet the upper bounds in rows 6–9 of Table 1.

Continuing the comparison between Ξ(G) and the generating graph of G from [10, 22], we note that in the non-abelian finite simple case, the diameter of the former graph may be larger than that of the latter, which is equal to 2 by [5, Theorem 1.2].  

Remark 1.3

It is easy to determine the connectedness and diameter of the non-commuting, non-generating graphs of certain infinite simple groups G. For instance, if G is a Tarski monster, then each nontrivial proper subgroup of G has a fixed prime order, and so each vertex of Ξ(G) is isolated. Additionally, it is a straightforward consequence of [1, Proposition 2.1] that diam(Ξ(G))=2 whenever G is not 2-generated. However, it is an open problem to investigate the connectedness and diameter of Ξ(G) when the infinite simple group G is 2-generated and Ξ(G) has an edge.

 
Remark 1.4

Our proof of Theorem 1.1 involves Proposition 3.10, which corrects [20, Theorem 1.1] on the intersection graph (defined in Section 2.1), in the case of the monster group. Additionally, our proof uses Theorem 4.5, on subspace stabilizers in simple unitary groups. In [21, Ch. 3], we prove more general versions of that theorem to obtain lower bounds for the base sizes of primitive subspace actions of almost simple classical groups.

We also note that our proof of Theorem 1.1(i) relies on the classification of finite simple groups. In Section 2.2, we prove the following special case that does not depend on the classification.  

Proposition 1.5

Let G be a non-abelian finite simple group, and suppose that every maximal subgroup of G has even order. Then Ξ(G) is connected with diameter at most 5.

In fact, it follows from the classification (see Theorem 2.6 and Remark 2.7) that in addition to a few sporadic groups, the finite simple groups containing maximal subgroups of odd order belong to only three infinite families. Thus the above proposition serves as a useful reduction for the proof of Theorem 1.1(i).

Now, if G/Z(G) is simple, then Z(G)x is a vertex of Ξ(G/Z(G)) whenever xGZ(G). Thus [22, Lemma 2.9(ii)] yields the following corollary of Theorem 1.1(i).  

Corollary 1.6

Let G be a group such that G/Z(G) is finite and simple. Then Ξ(G) is connected, with diam(Ξ(G))diam(Ξ(G/Z(G)))5.

If instead G/Z(G) is finite and almost simple but not simple, then we observe from [22, Theorems 1.2–1.3 and Lemma 6.5] that Ξ(G) is connected with diameter 2 or 3. Magma computations show that diam(Ξ(S5))=2 and diam(Ξ(S6))=3, and so both possibilities occur.

Finally, as a consequence of Theorem 1.1, we will prove the following analogue of that theorem for the non-generating graph Γ(G) of a non-abelian finite simple group G. Recall that this graph is the complement of the generating graph of G. Hence its vertex set is G{1}, and vertices x and y are adjacent if and only if x,yG. Note that Ξ(G) is a spanning subgraph of Γ(G) (since Z(G)=1). As above, we write B to denote the baby monster group and M to denote the monster group. Similar to Theorem 1.1, we incorporate new information on the maximal subgroups of M (compared to [21, Theorem 6.5.4]) and its intersection graph.  

Theorem 1.7

Let G be a non-abelian finite simple group. The non-generating graph Γ(G) is connected with diameter 4 or 5 if G=M, at most 4 if GM and at most 3 if every maximal subgroup of G has even order. Furthermore, if G{B,PSU(7,2)}, then diam(Γ(G))=4.

An earlier version of this theorem was communicated by this paper’s author to Peter Cameron, and appears as a remark after Proposition 12.3 in his paper [9]. That version of the theorem is a key component of the proof of [38, Theorem 1] (see the proof of [38, Proposition 10]), which investigates the connectedness and diameter of Γ(G) for a finite group G. This graph is closely related to the soluble graph of G, recently studied in [8]; see Remark 7 in that paper.

The remainder of this paper is structured as follows. The focus of Section 2 is on preliminary results and includes a comment on our computational methods and a proof of Proposition 1.5. In Section 3, we prove the alternating, sporadic and exceptional cases of Theorem 1.1. We consider matrices and stabilizers of subspaces in linear and unitary groups in Section 4 and then apply our results in Section 5 to prove Theorem 1.1 for those groups. Additionally, Section 5 concludes with a proof of Theorem 1.7.

2. PRELIMINARIES

Here, we present preliminary results on the non-commuting, non-generating graph of a group and on maximal subgroups of finite simple groups; in particular, we prove Proposition 1.5.

2.1. The non-commuting, non-generating graph

In this subsection, G is an arbitrary group, unless specified otherwise. Given vertices x and y of a graph Γ, we denote the distance in Γ between x and y by d(x,y), and if x and y are adjacent, then we write xy.  

Lemma 2.1
([10, Corollary 6])

Let H be a proper non-abelian subgroup of G. Then the induced subgraph of Ξ(G) corresponding to HZ(H) is connected with diameter 2.

 
Proposition 2.2

Suppose that G is finite and insoluble. Then Ξ(G) has no isolated vertices. Equivalently, each element of GZ(G) is a non-central element of some maximal subgroup.

 
Proof.

It is clear from the definition of Ξ(G) that a vertex xΞ(G) is isolated if and only if x lies in a unique maximal subgroup M of G and xZ(M). Since G is finite, it follows from [22, Proposition 2.6] that each isolated vertex of Ξ(G) lies in an abelian maximal subgroup of G. However, each maximal subgroup of the finite insoluble group G is non-abelian [29].

Our next result highlights a connection between Ξ(G) and the intersection graphΔG of G. This latter graph, introduced in [14], has vertex set the proper nontrivial subgroups of G, with subgroups adjacent if and only if they intersect nontrivially.  

Lemma 2.3

Let Γ be equal to Ξ(G) or the non-generating graph of G. Additionally, suppose that Z(G)=1 and that Γ is connected with a finite diameter. Then diam(Γ)diam(ΔG)1.

 
Proof.

By Propositions 12.1 and 2.3 of [9] (which assume that G is finite, but whose proofs also hold for infinite groups), the diameters of ΔG and the non-generating graph of G differ by at most one. As Z(G) is trivial, Ξ(G) is a spanning subgraph of the non-generating graph of G, and so the diameter of the former is at least the diameter of the latter.

We conclude this subsection with a brief description of our computational methods for determining diam(Ξ(G)) in Magma (for sufficiently small G). Let X be a set of representatives of generators for the non-central cyclic subgroups of G, and for each xX, let Yx be a set of representatives for the elements of X up to conjugacy in CG(x). We also obtain a set Xʹ from X by identifying elements if they generate conjugate cyclic subgroups. It is easy to see that Aut(Ξ(G)) contains Aut(G) and that diam(Ξ(G))=max{2,diam(Γ)}, where Γ is the graph obtained from Ξ(G) by identifying vertices that generate identical cyclic subgroups. We obtain diam(Ξ(G)) by calculating the maximum distance in Γ between the elements in each pair (x, y) with xX and yYx, via computations involving centralizers of elements and intersections of maximal subgroups.

The Magma code used in these calculations is contained in the file diam_nc_ng in [19]; see also [21, Appendix A] for further information. When using this code to calculate diam(Ξ(G)) in this paper, we always represent G as a permutation group of minimal degree. We note that the time and memory requirements for running this code are negligible for sufficiently small groups (such as A5), but are in the order of a few days and a few gigabytes for the largest groups investigated in this paper via these methods, namely, G2(3), PSL(4,3) and M23.

2.2. Maximal subgroups of non-abelian finite simple groups

Assume throughout this subsection that G is non-abelian, finite and simple.  

Lemma 2.4

Let L and M be maximal subgroups of G, with |L| even.

  • If Z(L) contains an involution a, then LZ(L) contains a G-conjugate of a.

  • L contains an involution that does not lie in Z(L)Z(M).

  • Suppose that |M| is even, and let a be an involution of LM. Then MZ(M) contains an involution b that does not commute with a.

  • Suppose that LM lies in Z(M) and contains an involution f. Additionally, let uLM. Then LM contains an involution fʹ with [u,f]1.

 
Proof.

(i) Suppose that Z(L) contains an involution a, so that CG(a)=L. By [25, Corollary 1], there exists gG such that aagCG(a)=L. In fact, gL, as aZ(L). Thus LLg=CG(ag), and so the involution ag is non-central in L.

(ii) We will prove that L contains an involution sZ(L)Z(M). By (i), there exists an involution yLZ(L). If yZ(M), then we can set s = y. If instead yZ(M) (so that LM), then CG(y)=M. Additionally, since Z(L) and LM are proper subgroups of L, there exists hL(Z(L)M), and yh is a non-central involution of L. Moreover, MMh=CG(yh). Therefore, yhZ(M), and so we can set s=yh.

(iii) Suppose first that |Z(M)| is odd, let S be the set of involutions of M and let QMM be the subgroup generated by S. If the involution aLM commutes with each element of S, then aCG(QM), and so QMM,a=G, contradicting the simplicity of G. Thus there exists rSCG(a). As SZ(M)=1, we can set b = r.

If instead Z(M) contains an involution z, then by applying (i) to M, we deduce that MZ(M) contains an involution c. If [a,c]1, then we can set b = c. Otherwise, as aM=CG(z), we see that [a,zc]=[a,z]c1. Thus we may set b to be the involution zcMZ(M).

(iv) Let s be the involution from (ii), so that sL and sLMZ(M). If [u,s]1, then we can set f=s. Assume therefore that [u,s]=1. Since CG(f)=M and similarly CG(fs)=MsM, we obtain fsLZ(M), and hence fsLM. If fsCG(u), then CL(u) contains s,fs, which contains f. This contradicts uM=CG(f), and so we can set f=fs.

The following lemma is one of the main ingredients in the proof of Theorem 1.1.  

Lemma 2.5

Let L and M be maximal subgroups of G of even order, and let xLZ(L) and yMZ(M). Then d(x,y)5, as vertices of Ξ(G). Moreover, if L contains an involution a such that d(x,a)1, then d(x,y)4.

 
Proof.

We will use Lemma 2.1 several times in this proof without further reference. First, we may assume that LM, as otherwise d(x,y)2. Let C be the set of involutions in LZ(L). Lemma 2.4(ii) shows that there exists cCZ(M). Notice that d(x,c)2. If cMZ(M), then d(c,y)2, and so d(x,y)d(x,c)+d(c,y)4. If instead cGM, then Lemma 2.4(iii) yields an involution bMZ(M) such that [c,b]1. As c,b is dihedral, cb. Since d(b,y)2, we conclude in this case that d(x,y)5, with d(x,y)4 if d(x,c)1.

Now, let A be the set of involutions aC with d(x,a)1. By the previous paragraph, it remains to consider the case where C(GM)Z(M) and AZ(M). Note that cCM in this case.

Let aAZ(M), so that CG(a)=M. Then ac, and since cA, it follows from the definition of A that ax and ax. Thus xM. Lemma 2.4(iv) (with f = a and u = x) now shows that LM⩽̸Z(M), as otherwise LM would contain an involution in A. Hence LM contains an element not centralizing M and the element a not centralizing x. As no group is the union of two proper subgroups, there exists hLM centralizing neither M nor x. Thus d(h,y)2 and xh, and therefore d(x,y)3.

 
Proof of Proposition 1.5

For each xΞ(G), Proposition 2.2 gives xKZ(K) for some maximal subgroup K of G (of even order). Thus diam(Ξ(G))5 by Lemma 2.5.

As the proofs of [25, Corollary 1] and Proposition 2.2 are independent of the classification of finite simple groups, so are the proofs of Lemmas 2.4 and 2.5 and Proposition 1.5.

To prove Theorem 1.1(i), it remains to consider the non-abelian finite simple groups that contain maximal subgroups of odd order.  

Theorem 2.6
([35, Theorem 2])

Suppose that G is not isomorphic to the monster group. Then G has a maximal subgroup of odd order if and only if G is isomorphic to one of the following:

  • Ap, with p being a prime, p3(mod4) and p{7,11,23};

  • PSL(n,q), with n being a prime, (n,q)(3,4) and q3(mod4) if n = 2;

  • PSU(n,q), with n an odd prime and (n,q){(3,3),(3,5),(5,2)}; or

  • the Mathieu group M23, the Thompson group Th or the baby monster group B.

 
Remark 2.7

Let G be the monster group. By [35, Theorem 2], any maximal subgroup of G of odd order has shape 59:29 or 71:35. It was claimed in [31, 32] that maximal subgroups of G isomorphic to PSL(2,59) and PSL(2,71) were constructed computationally (using data that were not made publicly available). This suggested that no maximal subgroup of shape 59:29 or 71:35 exists. However, a brand new computational paper [17] shows (via publicly available code) that, up to conjugacy, G has unique maximal subgroups of shape 59:29 and PSL(2,71), but no subgroup isomorphic to PSL(2,59). Hence, up to conjugacy, 59:29 is the unique maximal subgroup of G of odd order.

3. ALTERNATING, SPORADIC AND EXCEPTIONAL SIMPLE GROUPS

In this section, we prove the alternating, sporadic and exceptional cases of Theorem 1.1. Note that we consider the Tits group 2F4(2) together with the sporadic groups.

3.1. Alternating groups

Let G be the alternating group An of degree n5. We will prove the following result by considering the standard action of G on the set Ω:={1,2,,n}.  

Theorem 3.1

The graph Ξ(G) of the alternating group G of degree n5 is connected with diameter at most 4 or at most 3 if n is even.

Recall that a derangement of G is an element with no fixed points.  

Lemma 3.2

Let x be a derangement of G such that x acts intransitively on Ω. In addition, let {α1,α2} and {β1,β2} be 2-subsets of distinct orbits of x on Ω. Finally, if x has exactly two orbits on Ω, then let g:=(α1,α2)(β1,β2), and otherwise, let g:=(α1,α2,β1). Then xg.

 
Proof.

If x has exactly two orbits on Ω, then at least one of these orbits has length at least three. Hence, in each case, xgx, and so [x,g]1. Additionally, x,g acts intransitively on Ω and is therefore a proper subgroup of G. Thus xg.

 
Lemma 3.3

Let x,yG{1}. Then d(x,y)4. If d(x,y)3, then at least one of x and y is a derangement, and if d(x,y)=4, then both are derangements.

 
Proof.

Let α and β be distinct points in Ω. The point stabilizer GαAn1 is a maximal subgroup of G. As Z(Gα)=1, it follows from Lemma 2.1 that d(g,h)2 for all g,hGα{1}. Additionally, H:=GαGβAn2 is maximal in each of Gα and Gβ. Let gGα and kGβ, such that g,kH. Then CH(g)<H, as otherwise H,g=Gα would centralize g. Similarly, CH(k)<H. Hence some element of H centralizes neither g nor k, and it follows that d(g,k)2.

To complete the proof, we will show that each derangement xG is adjacent in Ξ(G) to some non-derangement. This follows immediately from Lemma 3.2 if x is intransitive on Ω. Assume therefore that x is an n-cycle (α1,,αn), with αiΩ for each i, and n is odd. It suffices to prove that there exists tNGα1(x) with xtx; it will then follow that x,tNG(x)<G and [x,t]1, so that xt. For each n-cycle rx with rx, there exists an element sN(Sn)α1(x) such that xs=r. There are ϕ(n)1>1 of these n-cycles, where ϕ is Euler’s totient function. We deduce that there exist s,sN(Sn)α1(x) such that xs=x1 and xs=xi for some i{2,,n2}. Since xss=xix, we can choose tG{s,s,ss}.

We will write supp(g) to denote the support {αΩαgα} of an element gG.  

Lemma 3.4

Let x and y be derangements of G such that each of x and y acts intransitively on Ω. Then d(x,y)3.

 
Proof.

Let {α1,α2} and B:={β1,β2} be two subsets of distinct orbits of x on Ω, such that |α1x||β1x|, and let {γ1,γ2} and D:={δ1,δ2} be 2-subsets of distinct orbits of y on Ω. We may assume that α1=γ1. If n = 5, then each of x and y has exactly two orbits on Ω.

Case (a): x and y each have exactly two orbits on Ω. We see from Lemma 3.2 that xa:=(α1,α2)(β1,β2) and yb:=(α1,γ2)(δ1,δ2). In particular, if a = b, then d(x,y)2.

Assume therefore that ab, and hence either α2γ2 or BD. If [a,b]=1, then either γ2=α2 and BD=; or {γ2}D={α2}B. As |α1x|3, we can repeat the argument with α2 replaced by an element of α1x{α1,α2} to obtain [a,b]1. Since a,b is dihedral, (x,a,b,y) is a path in Ξ(G) and d(x,y)3.

Case (b): x and y each have at least three orbits on Ω. By Lemma 3.2, xa:=(α1,α2,β1) and yb:=(α1,γ2,δ1). Observe that t:=supp(a)supp(b)5<n, and so a,b<G. Hence if [a,b]1, then (x,a,b,y) is a path in Ξ(G), and so d(x,y)3. If instead [a,b]=1, then t = 3 and b{a,a1}, hence ay and d(x,y)2.

Case (c): Exactly one of x and y has exactly two orbits on Ω. We may assume that x has exactly two orbits on Ω. Lemma 3.2 shows that xa:=(α1,α2)(β1,β2) and yb:=(α1,γ2,δ1). Observe that [a,b]1 and a,b is intransitive. Hence ab and d(x,y)3.

 
Proof of Theorem 3.1

Lemma 3.3 shows that Ξ(G) is connected with diameter at most 4. Furthermore, if n is even, then each derangement of G generates a subgroup that acts intransitively on Ω. Hence Lemmas 3.3 and 3.4 yield diam(Ξ(G))3.

Magma calculations show that diam(Ξ(An))=2 whenever 5n10. We do not know if there exists n > 10 such that diam(Ξ(An))>2.

3.2. Sporadic groups

Let G be a sporadic finite simple group, or the Tits group 2F4(2). In this subsection, we implicitly use [12, 43] to determine information about maximal subgroups, conjugacy classes and centralizers in G. In particular, all maximal subgroups of the sporadic groups are listed in [43] (up to conjugacy), with the exceptions of the maximal subgroups of the monster group M isomorphic to PGL(2,13) and Aut(PSU(3,4)) constructed in [18] and the maximal subgroups isomorphic to 59:29 constructed in [17] (see the introductions of those last two papers for further details). Additionally, the subgroup PSL(2,59) listed in [43] is not in fact a subgroup of M; see Remark 2.7.  

Theorem 3.5

The graph Ξ(G) of the sporadic group or Tits group G is connected with diameter 4 or 5 if G=M and at most 4 otherwise. If G{M11,M12,M22,M23,J1,J2,B}, then diam(Ξ(G)) is given in Table 1.

A few of the proofs in this subsection, where specified, involve straightforward computations in GAP [24] that utilize the Character Table Library [4], with runtimes of at most a few seconds. This library contains the necessary information regarding all maximal subgroups of G, except for certain maximal subgroups of the monster group.  

Proposition 3.6

Each maximal subgroup of G has a centre of order at most 2.

 
Proof.

For each maximal subgroup M of M not included in the GAP Character Table Library, no element of M has a centralizer of order |M|, hence Z(M)=1. For each other maximal subgroup of each G, we verify the result using GAP.

In light of Lemma 2.5, it will be useful to determine neighbours of involutions in Ξ(G).  

Proposition 3.7

Let x be an element of G of order at least 3, and suppose that x lies in a maximal subgroup of G of even order. Then x is adjacent in Ξ(G) to some involution of G.

 
Proof.

It suffices to show that there exists a maximal subgroup of G that contains both x and an involution that does not centralize x. Let M be a maximal subgroup of G of even order, with xM. Since |x|>2, Proposition 3.6 shows that xZ(M). Let TM be the set of involutions of M. If |TM|>|CM(x)|, or if the normal subgroup TM of M is equal to M, then x does not centralize every involution of M. Note that this second condition holds whenever M is simple.

Using GAP, we see that if x does not lie in any simple maximal subgroup, nor in any maximal subgroup M with |TM|>|CM(x)|, then one of the following holds:

  • G=J1 and |x|{7,19};

  • G=Ly and |x|{37,67};

  • G=Co1 and |x|=3;

  • G=J4 and |x|{29,43}; or

  • G=Fi24 and |x|=29.

Our computations here involve the fusion of M-conjugacy classes in G. As such, when G=M, we consider only the set of maximal subgroups of M for which this fusion is stored in GAP (this is a subset of the set returned by the NamesOfFusionSources function). Similarly, when G=B, our computations exclude the maximal subgroups of shape (22×F4(2)):2, for which this fusion is not stored.

Now, in case (iii), x is an element of the G-conjugacy class labelled 3A in the atlas [12] and lies in a maximal subgroup K(A5×J2):2, which clearly satisfies TK=K. In all other cases, |x| is odd and CG(x)=x, and hence x does not commute with any involution of M.

 
Lemma 3.8

Suppose that G=B, and let R and M be maximal subgroups of G isomorphic to Fi23. Additionally, let sR with |s|=23. Then there exists fRM such that d(s,f)1.

 
Proof.

Let S:=s. If SM, then we can set f = s. Otherwise, SM=1, and since |R||M|>|G|, there exists f(MR){1}. As fS=CR(s), we observe that sf.

 
Lemma 3.9

Suppose that G=B, and let s,yG{1}, with |s|=23. Then d(s,y)3.

 
Proof.

The element s generates a Sylow 23-subgroup S of G and lies in maximal subgroups RFi23 and T21+22Co2. Each maximal subgroup of G that contains s and is not G-conjugate to R or T has shape 47:23. Moreover, CR(s)=S and (using Proposition 3.6 and the atlas) CG(s)=CT(s)=Z(T)×SC2×S. In addition, Proposition 2.2 shows that yLZ(L) for some maximal subgroup L of G. We will show that d(s,y)3 whenever L satisfies certain properties and that we may often choose L to satisfy those properties.

Suppose first that Z(L)>1 and |L||T|>|G|, so that LT1. We may assume that LT, else d(s,y)2 by Lemma 2.1. If there exists a non-identity element u that lies in (LT)(Z(L)Z(T)) then Z(L)Z(T)LT (if uZ(L), say, then uZ(G) implies that Z(T)LT, and then Z(T)Z(G) yields Z(L)LT). Note also that Z(L)Z(T) is a proper subset of Z(L),Z(T), since Z(L)Z(T)=1. Thus regardless of whether such an element u exists, LT contains an element cZ(L)Z(T). Moreover, any such c satisfies d(c,y)2 by Lemma 2.1. In particular, if SLT, then we can set c = s, and so d(s,y)2. Assume therefore that S⩽̸LT. We claim that sc. Indeed, if sc, then c centralizes s and hence lies in (Z(T)×S)(Z(T)S). Thus c=zsi for some i{1,2,,22}, where z is the unique involution of Z(T). This implies that c2=s2i=S, and so SLT, a contradiction. Therefore, sc and d(s,y)3.

Next, suppose that Z(L)=1, |L||R|>|G| and LR. Then LR1, while SL=1 (since no L satisfying the given assumptions has order divisible by |S|). As CR(s)=S, each bLR is adjacent to s in Ξ(G). Moreover, d(b,y)2 by Lemma 2.1, and so again d(s,y)3.

Assume now that |y|{25,47,55}. Using GAP, we see that there exists a maximal subgroup K of G containing y, such that either Z(K)>1 and |K||T|>|G|; or Z(K)=1, |K||R|>|G| and KR. Moreover, Proposition 3.6 and Lemma 2.4(i) show that KZ(K) contains a G-conjugate of y. Hence we can choose L to be a G-conjugate of K, and d(s,y)3 by the previous two paragraphs. To complete the proof, we will consider the remaining elements y.

Case (a): |y|=47. Here, CG(y)=y, and each maximal subgroup of G containing y has shape 47:23. As S is a Sylow subgroup of G, it follows that ys for some s1 in a G-conjugate Sʹ of S. To complete the proof in this case, we will show that d(s,s)2.

The element sʹ lies in a G-conjugate M of R. If s=s, then it is clear that d(s,s)=2. Otherwise, [s,s]1, and so if s or sʹ lies in RM, then d(s,s)=1. Suppose finally that s,sRM. We deduce from Lemma 3.8 that d(s,f)=1=d(s,f) for some f,fRM. Hence there exists tRM centralizing neither s nor sʹ, and (s,t,s) is a path in Ξ(G).

Case (b): |y|{25,55}. There is a unique G-conjugacy class of elements of order |y|, and so we may choose LHN:2 if |y|=25 or L(5:4)×(HS:2) if |y|=55. In either case, no involution of G has a centralizer of order |L|, and so Proposition 3.6 yields Z(L)=1.

First suppose that there exists g(LR){1}. Then gSZ(L)Z(R), as |S| does not divide |L| and Z(L)=Z(R)=1. Hence Lemma 2.1 yields d(s,g),d(g,y)2. Additionally, CG(y)=y, and so CG(y)CG(s)=1. Thus either sg or d(g,y)1, and d(s,y)3.

Suppose finally that LR=1. Each of G and R has a unique conjugacy class of elements of order 35, and L also contains elements of order 35. Hence there exists a G-conjugate M of R such that LM contains an element m of order 35, and my since CG(y)=y. Additionally, Lemma 3.8 shows that d(s,f)1 for some fRM. Note that fMm=MCM(m), since mL and LR=1. Thus fm and d(s,y)3.

Recall that the vertices of the intersection graph ΔG of G are the proper nontrivial subgroups of G, and the graph’s edges are the pairs of subgroups that intersect nontrivially. The result [20, Theorem 1.1] on the diameter of ΔG (for all non-abelian finite simple groups G) does not take into account the new information about M mentioned in Remark 2.7. We now correct that theorem in the case G=M; this will be useful in the proofs of Theorems 3.5 and 1.7.  

Proposition 3.10

Suppose that G=M. Then the intersection graph ΔG of G is connected with diameter 5 or 6.

 
Proof.

Let M1 and M2 be maximal subgroups of G, with |M1| being odd. By elementary arguments from the proof of [20, Theorem 1.1], it suffices to bound d(S1,S2), where S1 and S2 are proper nontrivial subgroups of M1 and M2, respectively, such that S1 lies in no maximal subgroup of even order. By Remark 2.7, M1 has shape 59:29. Let S and H be subgroups of M1 of order 59 and 29, respectively, so that H is a Sylow subgroup of G.

We observe from [12] that |NG(H)|=2436. Arguing almost exactly as in the B case of the proof of [20, Theorem 1.1], with the aid of computations involving the GAP Character Table Library, we determine that M1 is the unique maximal subgroup of G containing S; that the maximal subgroups of G containing H are precisely 84 conjugates of M1, three conjugates of PGL(2,29), and one conjugate of 3.Fi24; and that d(S,Sg)5 for some gG. In particular, we may assume that S1=S.

Now, let R be a maximal subgroup of even order containing H. If S2 is not conjugate to S, then M2 can be chosen to have even order. Observe that the subgroup D generated by an involution of R and an involution of M2 is dihedral. Hence S1M1RDM2S2, and we obtain d(S1,S2)5. If instead S2=Sx for some xG, then we similarly deduce that S1M1RD^RxM1xS2 for some dihedral group D^, and thus d(S1,S2)6. Therefore, diam(ΔG){5,6}.

 
Remark 3.11

Let G, S and M1 59:29 be as in the proof of Proposition 3.10, and let F be a maximal subgroup of G of shape 3.Fi24. It follows from the proof of Proposition 3.10 that diam(ΔG)=6 if and only if there exists xG such that d(S,Sx)=6. Since M1 is the unique neighbour of S1 in ΔG, this occurs if and only if d(M1,M1x)=4.

One possible obstruction to this occurring is the existence of elements w,yG such that M1FwFyM1x. Since H lies in a unique conjugate of F, and M1 contains 59 conjugates of H, exactly 59 conjugates of F are adjacent to M1 in ΔG. An elementary counting argument now shows that exactly α:=59|F|/|M1|2.6×1023 conjugates of M1 are adjacent to F. Additionally, by the proof of [37, Theorem 1.3], the number β of conjugates of F adjacent to F is at most γ:=|G:F|rR|rGF|2|rG|, where R is a set of representatives for the G-conjugacy classes of elements of prime order. We calculate using the GAP Character Table Library that γ4.2×1028. We therefore deduce that there are at most 59αγ6.5×1053 conjugates M1x of M1 such that M1FwFyM1x for some w,yG. This is unfortunately not a useful upper bound, as it is larger than the number |G:M1|4.7×1050 of conjugates of M1.

Determining a better estimate for β is difficult due to the large index of F in G. However, if a sufficiently small upper bound for β can be obtained, then the above argument (along with similar arguments for the other types of path between M1 and M1x of length at most three) would yield the existence of an element x such that d(M1,M1x)=4, so that diam(ΔG)=6. Otherwise, deducing the exact value of diam(ΔG) would require more detailed information about the neighbourhoods of conjugates of M1, F and the maximal subgroup PGL(2,29) of G. The large order of G is again a significant obstacle to obtaining this information.

 
Proof of Theorem 3.5

For each G{M11,M12,M22,M23,J1,J2}, we obtain diam(Ξ(G)) using Magma. In all other cases, for i{1,2}, let xi be an element of G{1} that lies in a maximal subgroup of even order, say Li. By Proposition 3.7, d(xi,ai)1 for some involution ai of G, and so (using Proposition 2.2 if |xi|=2) we may choose Li so that xiZ(Li). Lemma 2.5 now gives d(x1,x2)4. Hence diam(Ξ(G))4 if each maximal subgroup of G has even order.

It remains to consider the case where G has a maximal subgroup K of odd order. By Theorem 2.6 and Remark 2.7, G{Th,B,M}. Let g,hG{1} such that h lies in no maximal subgroup of even order.

First assume that G=Th. Then K 31:15. Let mK. If |m|=31, then m lies in a maximal subgroup of G of shape 25PSL(5,2), and otherwise |CG(m)| is even. Thus each element of G lies in a maximal subgroup of even order, and so the element h does not exist. Hence the first paragraph of the proof yields diam(Ξ(G))4.

Suppose next that G=B. In this case, K 47:23, |h|=47 and CG(h)=h. Thus hs for each sK of order 23. As d(s,g)3 by Lemma 3.9, we conclude that d(h,g)4. It follows that diam(Ξ(G))4. In addition, the intersection graph of G has diameter 5 [20, Theorem 1.1], and so Lemma 2.3 yields diam(Ξ(G))4. Therefore, diam(Ξ(G))=4.

Finally, assume that G=M, so that K 59:29, |h|=59 and CG(h)=h. Thus hs for each sK of order 29. Let R be a maximal subgroup of G of even order containing s, and note that s is a Sylow subgroup of G. If g is not conjugate to h, then, as in the first paragraph of the proof, d(g,a)1 for some involution a of G, there exists a maximal subgroup L of G such that g,aLZ(L). By Lemma 2.4(iii) and the fact that two involutions generate a dihedral group, R contains an involution b such that ab. Moreover, |CG(s)|=87, which implies that sb. Hence hsbag, and so d(h,g)4. If instead g=hj for some jG, then we similarly deduce the existence of involutions aʹ and bʹ such that hsbasjg, and so d(h,g)5. Therefore, diam(Ξ(G))5. As the intersection graph of G has diameter at least 5 by Proposition 3.10, Lemma 2.3 implies that diam(Ξ(G)){4,5}.

Determining the exact diameter of Ξ(M) (and similarly for the non-generating graph of M in Theorem 1.7) is difficult for reasons similar to those discussed in Remark 3.11.

3.3. Exceptional groups of Lie type

We now prove Theorem 1.1 in the case where G is an exceptional group of Lie type. In view of Proposition 1.5 and Theorem 2.6, it remains to prove the following. Here, q denotes a prime power, and we do not include 2G2(3)PSL(2,8); for that group, see Theorem 5.1.  

Theorem 3.12

Let G be a finite simple group. If GSz(q), or if q is odd and G{G2(q),2G2(q),3D4(q),F4(q),E8(q)}, then Ξ(G) is connected with diameter at most 4.

Given a group X, let QX be the (characteristic) subgroup generated by the involutions of X.  

Lemma 3.13

Let S:=SL(2,q), and let H be a quasisimple group containing a subgroup K, such that KS and Z(H)=Z(K). Additionally, let R be the central product SH, and let π be the natural epimorphism S×HR. If QRR, then q = 3, QR is a maximal subgroup of R that contains (H)π, and CR(QR)=Z(R).

 
Proof.

We observe that S~:=(S)π and K~:=(K)π are normal subgroups of J:=SK=(S×K)π and S~ and H~:=(H)π are normal subgroups of R, with S~S, K~K and H~H. Additionally, [H~,S~]=([H,S])π=1 and H~S~=R. Furthermore, the centres of R, J, H~, S~ and K~ all coincide, and this common centre Z of order (2,q1) is equal to H~S~. We divide the rest of the proof into two cases.

Case (a): q3. We first show that QJ=J. This is clear if q = 2, so assume otherwise. Considering K as a copy of SL(2,q), let L:={(A,B)S×KA2=B2=I2}, where I2 is the 2×2 identity matrix. The choices for A and B include the matrix

$(0aa10)$
for each aFq×. As S is quasisimple, and a quasisimple group’s proper normal subgroups are precisely its central subgroups, we see that LS×K projects onto each of S and K and that the kernel of each projection map is isomorphic to SK. Thus LS×K=S×K. Since the image under π of each element of L is an involution, it follows that QJ=J.

Now, since H~ is quasisimple and Z<K~H~, no proper normal subgroup of H~ contains K~. Hence no proper normal subgroup of R contains J. As J=QJQRR, we obtain QR=R.

Case (b): q = 3. Here, QJ is the image under π of the direct product of the normal Sylow 2-subgroups of S and K, which are isomorphic to the quaternion group Q8. Thus S~QJ is maximal in S~ and K~QJ>Z. As the normal subgroup H~QR of H~ contains K~QJ, it follows that H~QR. Since R=H~S~, we deduce that QR=H~S~QR=H~(S~QR).

If S~QR=S~, then QR=R. We will therefore assume that S~QR<S~. Then S~QR is the maximal subgroup S~QJ of S~, and QR=H~(S~QJ), which is maximal in H~S~=R. Since [H~,S~]=1, we obtain ZCR(QR)=CH~S~(H~(S~QJ))Z(H~)CS~(S~QJ). The centralizer CS~(S~QJ) is equal to Z=Z(H~)=Z(R), and hence so is CR(QR).

For an example where QRR, take H to be the quasisimple group SL(2,9).  

Proposition 3.14

Let H be a finite quasisimple group. If |Z(H)|=2, then assume that H/Z(H) is not isomorphic to A7 or to PSL(2,q) for any odd q. Then QH=H.

 
Proof.

This is a straightforward consequence of the main theorem of [27], which implies that there exists an involution aHZ(H) if |Z(H)| is even. Clearly, a also exists if |Z(H)| is odd. Each proper normal subgroup of the quasisimple group H lies in Z(H), and so QH=H.

 
Lemma 3.15

Let G be a simple group in {G2(q),2G2(q),3D4(q),F4(q),E8(q)} with q odd (so that q27 if G=2G2(q)), let a be an involution of G and let Y:=CG(a). Then CY(QY)=Z(Y)=a.

 
Proof.

We observe from [26, pp. 171–177] that Z(Y)=a, and

  • G2G2(q) and YC2×PSL(2,q);

  • GF4(q) and YSpin(9,q);

  • GE8(q) and YJ:2, where J:=(Spin+(16,q)/A), in which A is a central subgroup of Spin+(16,q) of order 2; or

  • Y=R:A, with |A|=2 and RSL(2,q)H, where (G, H) lies in the set
    Additionally, Z(Y)=Z(R), and A induces an outer automorphism on the image of H under the natural epimorphism π:SL(2,q)×HR.

We will divide the remainder of the proof into the above cases. In each case, since Z(Y)=a, it remains to show that CY(QY)=Z(Y). Note that this is certainly the case if QY=Y.

Cases (a)–(c). By applying Proposition 3.14 to Y in Case (b) and J in Case (c), we easily deduce that QY=Y in each case.

Case (d). If GG2(3), then YSO+(4,3) [42, p. 125], and QY=Y. Suppose therefore that GG2(3) and that QY<Y. Notice that QY contains QR,A=QR:A, and thus QR<R.

In order to apply Lemma 3.13, we will show that the quasisimple group H contains a subgroup K such that KSL(2,q) and Z(H)=Z(K). Such K clearly exists if there exists LH with Z(H)=Z(L) and LSL(2,qi) for some positive integer i. Hence this is the case if H{SL(2,q),SL(2,q3)}. If instead H=E7(q)sc, then H contains a subgroup LSL(2,q7), with Z(H)=Z(L) [28, p. 927, item (3)]. Finally, if H=Sp(6,q), then let V1, V2 and V3 be non-degenerate two-dimensional subspaces of the symplectic space V:=Fq6, such that V=V1V2V3. Then the intersection of the stabilizers in H of V1, V2 and V3 is isomorphic to Sp(2,q)3=SL(2,q)3. The diagonal subgroup of this direct product is isomorphic to SL(2,q) and has centre Z(H).

Lemma 3.13 now shows that QR is a maximal subgroup of R, that (H)πQR and that CR(QR)=Z(R). It follows from this first fact that the proper subgroup QY of Y is maximal and equal to QR:A. Since (H)πQRR=(SL(2,q))π(H)π and [(SL(2,q))π,(H)π]=1, each inner automorphism of QR acts on (H)π as an inner automorphism of (H)π. Thus the outer automorphism of (H)π induced by A does not extend to an inner automorphism of QR. Therefore, CY(QR)=CR(QR)=Z(R), which is equal to Z(Y) from the start of the proof. As Z(Y)CY(QY)CY(QR), we conclude that CY(QY)=Z(Y).

 
Proof of Theorem 3.12

Theorem 2.6 implies that each maximal subgroup of G has even order, and so it suffices by Proposition 2.2 and Lemma 2.5 to show that each non-involution xG{1} is adjacent in Ξ(G) to some involution. Let M be a maximal subgroup of G containing x. Assume first that GSz(q), so that q is odd, and let a be an involution of M. If xa, then we are done. Otherwise, xCG(a). By Lemma 3.15, a is the unique non-identity element of CG(a) that centralizes each involution of CG(a). Hence xa for some involution aCG(a).

Assume now that GSz(q). By Theorem 4 on page 122 of [41], CG(y) is nilpotent for each yG{1}. As the only non-abelian finite simple groups with nilpotent maximal subgroups are PSL(2,p) for various primes p (see [40, p. 183]), we deduce that Z(M)=1 and CM(x)<M. If MSz(q0), for some proper power q0 of 2 dividing q, then M=QM. It follows in this case that M contains an involution a not centralizing x, and so xa. If instead MSz(q0), then by Section 4, page 133 and Theorem 9 on page 137 of [41], M is conjugate in G to

  • the Frobenius group W:=S:Cq1, where S is a Sylow 2-subgroup of G;

  • the dihedral group X:=D2(q1); or

  • a Frobenius group Y±:=Cq±2q+1:C4.

If MX, then again QM=M and xa for some involution aM. If instead MY±, then let N be the normal subgroup Cq±2q+1 of M. As M is Frobenius, there exists a complement H of N with xH. No non-identity element of H centralizes x, and so x is adjacent in Ξ(G) to the unique involution of H.

Suppose finally that MW. If xS, then x centralizes no non-identity element of S, and so x is adjacent in Ξ(G) to each involution of S. Now, Z(S) consists of all elements of S of order at most 2, and each element of SZ(S) has order 4, by Lemma 1 on page 112 and Theorem 7 on page 133 of [41]. Additionally, G has a unique conjugacy class of cyclic subgroups of order 4 (see the final sentence of page 121 and Proposition 18 on page 126 of [41]), and so if the non-involution x lies in S, then it also lies in a maximal subgroup LY±. By the previous paragraph, x is adjacent in Ξ(G) to an involution of L.

Magma calculations show that diam(Ξ(Sz(8)))=3 and diam(Ξ(G2(3)))=2. We do not know if any finite simple exceptional group G satisfies diam(Ξ(G))>3. Observe also that the proof of Theorem 3.12 relies on useful properties of the involution centralizers in G (and of Sz(q) and its maximal subgroups). Indeed, in the q odd case, we have considered precisely those finite simple groups whose involutions are all centralized by maximal subgroups of maximal rank (see [26, Theorem 4.5.1] and [36]). The involution centralizers for the remaining exceptional groups have more complicated structures (see for example [2] and [26, Theorem 4.5.1]), so alternative methods would be necessary to prove an analogue of Theorem 3.12 for those groups.

4. Matrices and subspace stabilizers

We now consider various matrices and subspace stabilizers. The results here will be useful when proving Theorem 1.1 for linear and unitary groups. Let n be a positive integer, q a prime power and Ik the k × k identity matrix over Fq, for a positive integer k. Additionally, let Ei,j be the n × n matrix with (i, j) entry equal to 1 and all other entries equal to 0.

4.1. Matrices and subspace stabilizers in linear groups

([39, p. 148 & pp. 161–162])

 
Definition 4.1
Let f(x):=xnβnxn1β2xβ1 be a (monic) polynomial in Fq[x]. The companion matrix of f is the n × n matrix
where
$A:=(β1)$
,
$B:=(β2βn)$
and C(f)=A when n = 1. If f is irreducible over Fq, then for each positive integer k, the hypercompanion matrix of f k is the kn × kn matrix

Since a subgroup A of GL(n,q) is irreducible if and only if the characteristic polynomial χA of A is irreducible, the following lemma will shed light on certain irreducible subgroups.  

Lemma 4.2

Let aFq×. Then the set of irreducible factors in Fq[x] of the binomial xq1a is {x|a|bbFq×,b(q1)/|a|=a}.

 
Proof.

For each positive integer k, let ϕk be the homomorphism ssk of Fq×, and let Hk be the image of ϕk, of order (q1)/(k,q1). Additionally, let r:=|a| and t:=(q1)/r. Then Ht=a. Since |ker(ϕt)|=t, there are exactly t distinct roots b1,,btFq× of the binomial f(x):=xta. Hence f(x)=i=1t(xtbi), and xq1a=f(xr)=i=1t(xrbi).

It remains to show that the binomial xrbi is irreducible over Fq for each i. Since bit=a, we see that r=|a| divides |bi|. Additionally, if r is divisible by 4, then so is q − 1. It therefore suffices by [34, Theorems 3.3 and 3.35] to prove that r is coprime to u:=(q1)/|bi|.

Since |Hu|=|bi|, there exists ciFq× with ciu=bi, and hence ciut=a. Thus aHut, and so r divides |Hut|=(q1)/(ut,q1)=tr/(ut,tr)=r/(u,r). Hence (u,r)=1.

Recall that matrices A,BGL(n,q) are similar if and only if χA=χB.  

Lemma 4.3

Let ASL(n,q), and suppose that Aq21Z(SL(n,q)) and that A acts irreducibly on Fqn. Then n{1,2}. If, in addition, n = 2 and Aq1Z(SL(2,q)), then q3(mod4).

 
Proof.

Since Aq21Z(SL(n,q)), there exists aFq× such that Aq21aIn=0. Thus χA, which is irreducible over Fq, divides xq21aFq[x]. Furthermore, Fqn is the splitting field for χA over Fq, and A is similar to the companion matrix C(χA). To prove the required result, we will study χA as a polynomial in Fq2[x]. By Lemma 4.2, the set of irreducible factors in Fq2[x] of xq21a is {x|a|bbFq2×,b(q21)/|a|=a}.

If n is odd, then χA is irreducible over Fq2. Thus n=|a|, and χA=xnb for some bFq×, with b(q21)/n=a. Since det(C(xnb))=b, we deduce that b = 1. Hence a = 1 and n = 1.

If instead n is even, then each irreducible factor of χA in Fq2 has degree n/2. In particular, |a|=n/2, and there exist b1,b2Fq2× such that
with b1(q21)/(n/2)=b2(q21)/(n/2)=a and b1b2,b1+b2Fq. Here, det(C(χA))=b1b2, and thus b2=b11. Therefore, a=a1, and hence n/2=|a|{1,2}. In particular, if q is even, then aFq× has odd order, and n = 2. Otherwise, arguing as in [11, pp. 569–570] shows that the irreducible polynomial xn(b1+b2)xn/2+1Fq[x] cannot have degree 4, and again n = 2.

Finally, if n = 2 and Aq1Z(SL(2,q)), then Aq1cI2=0 for some c{1,1}. Hence the irreducible polynomial χA divides xq1c. It follows from Lemma 4.2 and the previous paragraph that χA(x)=x2+1, which is irreducible if and only if q3(mod4).

Let H:=SL(n,q), with n2 and q4 if n = 2, so that H/Z(H) is simple. Additionally, let V:=Fqn be the natural module for H and let HX be the stabilizer in H of a subspace X of V.  

Lemma 4.4

Let X and Y be distinct one-dimensional subspaces of V. Then:

  • no commutator in HX is a non-identity scalar matrix;

  • CH(HX)=Z(H); and

  • $C_H(H_X \cap H_Y) = {HXHY,if n=2,or n=3\,andq=2,Z(H),otherwise.$

 
Proof.

Up to conjugacy by a fixed element of H, the stabilizer HX consists of the matrices in H with a first row of the form

$(λ1000)$
⁠, with λ1Fq×, and HY is the set of matrices with a second row of the form
$(0λ200)$
, with λ2Fq×. The restriction of a matrix in HX to its (1, 1) entry is a homomorphism from HX to the abelian group Fq×, and its kernel clearly contains all commutators in HX. Thus (i) holds. We split the proof of (ii) and (iii) into two cases. In each case, let WCH(HXHY).

Case (a): n = 2. Let ω be a primitive element of Fq, A:=diag(ω,ω1)HXHY and D:=I2+E2,1HX. The equality AW = WA gives WHXHY, which is abelian. Hence we obtain (iii). If WCH(X), then DW = WD, and we deduce that WZ(H), hence (ii).

Case (b): n3. It is easy to check the result if n = 3 and q = 2, so assume otherwise. If n = 3, then let bFq×{1} and B:=diag(1,b,b1),B:=diag(b,1,b1)HXHY. We deduce from the equalities BW = WB and BW=WB that W is diagonal. If instead n4, then let r and s be distinct integers with 3rn and 1sn. Then Cr,s:=In+Er,sHXHY, and the equalities Cr,sW=WCr,s (for all r and s) imply that W is again diagonal.

In each case, let FHXHY be the matrix obtained from In by setting all entries in the final row to 1. As W is diagonal, the equality WF = FW yields WZ(H). Thus we have proved (iii), and (ii) follows immediately.

4.2. Subspace stabilizers in unitary groups

Throughout this subsection, let n3, with q3 if n = 3, so that G:=PSU(n,q) is simple. Additionally, let V be the natural module Fq2n for SU(n,q), equipped with the unitary form with Gram matrix In. Then a matrix ASL(n,q2) lies in SU(n,q) if and only if AAσT=In, where σ is the automorphism ααq of Fq2. The following theorem will be useful when considering Ξ(G). Here, G(Δ) denotes the pointwise stabilizer in G of a set Δ of subspaces of V.  

Theorem 4.5

Let Δ be a set of k-dimensional subspaces of V.

  • Suppose that |Δ|<n/k. If (q+1)n, or if k(n1), then G(Δ)1.

  • Suppose that |Δ|<n/k1. Then G(Δ)1.

An analogue of the following lemma holds for certain classical algebraic groups; see [7, Section 4.1].  

Lemma 4.6

Let W be an (n1)-dimensional subspace of V. Then SU(n,q) contains a non-scalar matrix that acts as a scalar on W if and only if W is totally singular or (q+1)n.

 
Proof.

First, SU(n,q) has two orbits on the set of one-dimensional subspaces of V: the non-degenerate subspaces and the totally singular subspaces (cf. Witt’s Lemma). Thus it suffices to consider two choices for W, one for each possible orbit containing W. Let {e1,,en} be the basis for V corresponding to the Gram matrix In, and let ω be a primitive element of Fq2.

Case (a): W is non-degenerate. We may assume that W=e1, so that W=e2,e3,,en. Any matrix in SL(n,q2) that acts as a scalar on W also stabilizes W and hence is equal to Bα:=diag(α(n1),α,,α) for some αFq2×. Moreover, BαSU(n,q) if and only if αq+1=1.

Let λ:=ωq1. Observe that if (q+1)n, then Bλ is a non-scalar matrix in SU(n,q) that acts as λ on W. If instead q+1n, then any scalar αFq2× satisfying αq+1=1 also satisfies α(n1)=α. Therefore, in this case, no non-scalar matrix in SU(n,q) acts as a scalar on W.

Case (b): W is totally singular. If q is even, then let γ:=ωq+1 and δ:=1, and otherwise, let γ:=ω and δ:=ω(q1)/2. Then δq+1=1 for all q, and so δe1+e2 is singular. Hence we may assume that W=δe1+e2, so that W=δe1+e2,e3,,en. The direct sum of

$\left( 1+γδγγδ21γδ \right)$
and In2 is a non-scalar matrix in SU(n,q) that acts trivially on W.

 
Corollary 4.7

Let U be an (n2)-dimensional subspace of V. Then SU(n,q) contains a non-scalar matrix that acts as a scalar on U.

 
Proof.

Observe that U contains a one-dimensional totally singular subspace X. By Lemma 4.6, some non-scalar matrix in SU(n,q) acts as a scalar on X and hence on UX.

 
Proof of Theorem 4.5

If |Δ|<n/k, then the subspace Δ of V lies in an (n1)-dimensional subspace W. Lemma 4.6 shows that if (q+1)n, then SU(n,q) contains a non-scalar matrix that acts as a scalar on W and hence on Δ. Therefore, G(Δ)1.

Finally, assume that |Δ|<n/k1 or that |Δ|=n/k1 and k(n1). Then Δ lies in an (n2)-dimensional subspace U of V. By Corollary 4.7, SU(n,q) contains a non-scalar matrix that acts as a scalar on U and hence on Δ. We again conclude that G(Δ)1.

5. LINEAR AND UNITARY GROUPS AND THE NON-GENERATING GRAPH

In this section, we complete the proof of Theorem 1.1 by showing that it holds for linear and unitary groups. In addition, we prove Theorem 1.7. Define n, q, Ik and Ei,j as in Section 4, and let H{SL(n,q),SU(n,q)}, Z:=Z(H) and G:=H/Z. We will largely work in the group H, appealing to the following consequence of the definition of Ξ(G): for A,BHZ, the vertices ZA and ZB of Ξ(G) are adjacent if and only if [A,B]Z and A,B,Z<H.

5.1. Linear groups

Assume that H=SL(n,q), with n2 and q4 if n = 2, so that G is simple.  

Theorem 5.1

Let G be the linear group PSL(n,q). Then diam(Ξ(G))4. Moreover, if n = 2, then diam(Ξ(G))=2 if q is even or q9, and diam(Ξ(G))=3 otherwise.

Let U be the set of one-dimensional subspaces of the natural module V:=Fqn for H. We proceed by considering paths in Ξ(H) containing elements that stabilize subspaces in U.  

Lemma 5.2

Let X,YU, KHXZ and LHYZ. Then d(ZK,ZL)2.

 
Proof.

We may assume that KL. We claim that there exists a path (U1,,Uj) in Ξ(H), where j3, U1=K, Uj=L and U2HXHY. Since Z(HX)=Z(HY)=Z by Lemma 4.4(ii), the claim follows from Lemma 2.1 if K or L lies in HXHY. Otherwise, Lemma 4.4(iii) implies that CHXHY(K) and CHXHY(L) are proper subgroups of HXHY. Thus there exists an MHXHY centralizing neither K nor L, and we can set U2=M.

Now, [Ui,Ui+1]1 for all i{1,,j1}, and so [Ui,Ui+1]Z by Lemma 4.4(i). Moreover, Ui,Ui+1,Z lies in HX or HY. Hence (ZU1,,ZUj) is a path in Ξ(G).

 
Lemma 5.3

Suppose that n > 2 or q3(mod4), and let AH such that A stabilizes no subspace in U. Then there exist XU and KHX such that ZAZK.

 
Proof.

We divide the proof into two cases.

Case (a): A acts reducibly on V. It is clear that n > 2. The matrix A is similar to a block matrix whose final row of blocks is

$(00R)$
⁠, with R being a hypercompanion matrix and a proper submatrix of A (see [39, Theorem 8.10]). Since Aut(G)Aut(Ξ(G)), we may assume without loss of generality that A is this block matrix.

Let K:=In+E2,1. Then KHX, where XU is spanned by

$(100) \in V$
⁠. Each matrix in Z{K} can be viewed as a block matrix, with a final row
$(00S)$
of blocks, for some nonzero block S with the same dimensions as R. Hence A,K,Z<H.

It remains to prove that [A,K]Z. For a contradiction, suppose otherwise, so that (K1)AK=cIn for some cFq×. Then (K1)A=cK1. As K has eigenvalue 1 with algebraic multiplicity n, while cK−1 has eigenvalue c, we deduce that [A,K]=1. However, (AK)11=1 and (KA)11=0, a contradiction.

Case (b): A acts irreducibly on V. By [15, Proposition 3.6.1], A lies in a Singer subgroup S of GL(n,q), that is, a cyclic subgroup of order qn1. Additionally, N:=NGL(n,q)(S) is the group S:B of order n(qn1), where BGL(n,q) is similar to the companion matrix C(xn1) and satisfies FB=Fq for each FS (see [33, pp. 187–188] and [30, p. 497]). As C(xn1) stabilizes the subspace in U spanned by the all-ones vector, B also stabilizes some XU. Note that det(B)=det(C(xn1)) is equal to 1 if q is even or n is odd, and −1 otherwise. Additionally, N is soluble, and so its subgroup A,B,Z is proper in H. Furthermore, [A,B]=A1AB=Aq1. Since Aq21Aq1, Lemma 4.3 and our assumptions on n and q yield [A,B]Z.

It again suffices to find a matrix KHX with [A,K]Z and A,K,Z<H. If n is odd or q is even, then set K = B. If instead n = 2 and q1(mod4), then set K=αB for some square root αFq of −1. Finally, if q is odd and n is even and greater than 2, then B2HX. Additionally, [A,B2]=Aq21, which does not lie in Z by Lemma 4.3. Thus we set K=B2.

 
Proof of Theorem 5.1

Let x,yG{1}. We will bound d(x,y). By Lemma 5.2, d(x,y)2 if x and y stabilize one-dimensional subspaces of V. Since diam(Ξ(G))2, we may assume without loss of generality that x stabilizes no such subspace. If n3, then Lemma 5.3 shows that xs for some sGX with XU. Similarly, d(y,t)1 for some tGY with YU. Hence d(x,y)d(x,s)+d(s,t)+d(t,y)4, as required.

Suppose from now on that n = 2. Magma computations show that diam(Ξ(G))=2 when q is odd and at most 9, so we will assume that q is even or at least 11. Let p be the prime dividing q, and let r:=(2,q1). In what follows, all information about elements and subgroups of G is from [16, Ch. XII], except where stated otherwise. We first note that |G|=q(q21)/r and that each non-identity element of G lies in either a unique cyclic subgroup of order (q1)/r, a unique cyclic subgroup of order (q+1)/r, or a unique Sylow p-subgroup isomorphic to the elementary abelian group E of order q. Moreover, all cyclic subgroups of G of a given order (q±1)/r are conjugate. In addition, each element of order dividing q or (q1)/r lies in a parabolic maximal subgroup of shape E:q1r, which is the stabilizer in G of a subspace in U.

It now follows that |x| divides (q+1)/r. The unique cyclic subgroup of order (q+1)/r containing x lies in a unique maximal subgroup LD2(q+1)/r. We divide the remainder of the proof into two cases.

Case (a): q is even. Suppose first that |y| does not divide q − 1. All involutions of G are conjugate, as are all cyclic subgroups of order q + 1, and so y lies in a G-conjugate M of L. As Z(L)=1, Lemma 2.1 yields d(x,y)2 if L = M. If instead LM, then LM contains an involution a [23, Lemma 2.3], and it is clear that CL(a)=a=CM(a). Hence xa and d(a,y)1, and so d(x,y)2.

Next, suppose that |y| divides q − 1. Then y lies in a parabolic subgroup P of G of shape E:(q1). As |P||L|=2|G|, there exists an involution bPL. The centralizer of b in G is the Sylow 2-subgroup of P, and thus xby. Therefore, diam(Ξ(G))2.

Case (b): q is odd. There exist G-conjugates J and K of L with JK=1 [23, Lemma 2.2]. As above, J is the unique maximal subgroup of G containing an element s of order (q+1)/2 and K is the unique maximal subgroup containing a conjugate t of s. Neither s nor any neighbour of s in Ξ(G) is a neighbour of t, and so d(s,t)3. Hence it remains to show that d(x,y)3.

Suppose first that |y| divides (q+1)/2. Notice that Z(L)=1 if q1(mod4), else |Z(L)|=2. Additionally, since all cyclic subgroups of G of a given order (q±1)/2 are conjugate, so are all involutions of G. It follows that there exist G-conjugates R and S of L with xRZ(R) and ySZ(S). If R = S, then d(x,y)2 by Lemma 2.1. Otherwise, as the dihedral group R is generated by a pair of non-commuting involutions, there exists an involution aR(Z(R)S). In addition, Lemma 2.4(iii) yields an involution bSZ(S) with [a,b]1. Since a,b is dihedral, with CR(a)=Z(R)×a and CS(b)=Z(S)×b, we deduce that (x,a,b,y) contains a path in Ξ(G) from x to y. Hence d(x,y)3.

Assume from now on that |y| does not divide (q+1)/2, so that y lies in a parabolic subgroup P of shape E:q12, with P=GX for some XU. If q1(mod4), then Lemma 5.3 yields xk for some kGY with YU. Since d(k,y)2 by Lemma 5.2, we obtain d(x,y)3.

Suppose therefore that q3(mod4), and let c be an element of P of order (q1)/2. Then c is the pointwise stabilizer G(X,Yc) for some YcU{X}, and Nc:=NG(c) is dihedral of order q − 1 and equal to the setwise stabilizer G{X,Yc}. It is also clear that cf for each of the (q1)/2 involutions fNc, and in fact yf if yc, since |y|2. In addition, CG(c)=c, and so if yc, then ycf. Now, let cʹ be an element of Pc of order |c|, so that YcYc. Then NcNc=G{X,Yc}G{X,Yc}=G(X,Yc)G(X,Yc)=cc=1. Since there are q cyclic subgroups of order |c| in P, it follows that y has distance at most two from at least m:=q(q1)/2 involutions of G. This in fact accounts for all involutions of G, since the centralizer in G of an involution is conjugate to L, which has order |G|/m. Hence if |x|=2, then d(x,y)2. Otherwise, x is adjacent to each involution in LZ(L), and so d(x,y)3, completing the proof.

Using Magma, we see that Ξ(G) has diameter 2 when G{PSL(3,4),PSL(4,2)} and diameter 3 when G{PSL(3,3),PSL(3,5),PSL(3,7),PSL(4,3)}. Note that the diameter of Ξ(PSL(3,4)) was incorrectly reported as 3 in [21].

5.2. Unitary groups

Assume that H=SU(n,q), with n3 and q3 if n = 3, so that G is simple. In fact, by Proposition 1.5 and Theorem 2.6, we may assume that n is prime.  

Theorem 5.4

Let G be the unitary group PSU(n,q), with n being an odd prime.

  • Ξ(G) is connected with diameter at most 5.

  • Suppose that n = 7 and q = 2. Then Ξ(G) is connected with diameter 4.

Let V be the natural module Fq2n for H=SU(n,q), and fix a basis {e1,,en} for V so that the unitary form on V has Gram matrix In.  

Lemma 5.5

Suppose that n = 3, let ω be a primitive element of Fq2 and let λ:=ωq1. Let B1:=diag(λ2,λ,λ), B2:=diag(λ,λ2,λ) and B3:=diag(λ,λ,λ2). Additionally, let A1 and A2 be H-conjugates of B1.

  • BiHZ for each i.

  • B2 and B3 are each conjugate to B1 in H.

  • A1,A2 stabilizes a one-dimensional subspace of V.

  • If [Bi,A1]=1 for all i, then A1{B1,B2,B3}.

 
Proof.

Since λq=λ1 and det(Bi)=1, we see that BiH. Additionally, ω3(q1)1, and so BiZ, hence (i). For (ii), note that the matrices

$\left( 010100001 \right)$
and
$\left( 001010100 \right)$
lie in H and conjugate B1 to B2 and B3, respectively.

To prove (iii), we may assume that A1=B1. Let FH such that A2=B1F, and note that B1 and A2 act as λ on the subspaces E:=e2,e3 and EF of V, respectively. Hence B1,A2 stabilizes each subspace of EEF, which has positive dimension since dim(V)=3.

Finally, CH({B1,B2,B3}) consists of diagonal matrices, and all H-conjugates of B1 have the same characteristic polynomial as B1. Since B1, B2 and B3 are the only diagonal matrices in GL(n,q2) with this characteristic polynomial, we obtain (iv).

 
Lemma 5.6

Suppose that n is an odd prime.

  • The companion matrix C(xn1) is an element of H.

  • Let R be an element of H that is conjugate to C(xn1). Then ZR is adjacent in Ξ(G) to an involution of G.

  • Let R and S be H-conjugates of the companion matrix C(xn1). Then d(ZR,ZS) is at most 3 if n = 3 or at most 2 otherwise.

 
Proof.

(i) This holds since C(xn1)1=C(xn1)T and det(C(xn1))=1.

(ii) We may assume that R=C(xn1). If q is odd, then let A be the direct sum of the matrices In2 and I2, so that the (1, 1) and (n, n) entries of [R,A] are equal to −1 and 1, respectively. If instead q is even, then let AH be the direct sum of the matrices In2 and

$(0110)$
⁠, so that the (1, 1) entry of [R,A] is equal to 0. In either case, [R,A]Z. Moreover, R and A are monomial, and so R,A,Z<H. Thus ZRZA. As A is an involution, so is ZA.

(iii) We may assume that R=C(xn1) and d(ZR,ZS)>1. Then R stabilizes the one-dimensional subspace X of V spanned by the all-ones vector. Let FH such that S=RF, and let L:=HX, so that RL and SLF. Additionally, let L0:=He1. It is easy to show that CL0(R)=Z, and thus CL0F(S)=Z. Furthermore, as ZL and L < H, it follows from Lemma 4.4(i) that if [R,A]1 for some AL, then ZRZA. Similarly, if [S,D]1 for some DLF, then ZSZD. We split the remainder of the proof into three cases.

Case (a): n7, or n = 5 and q4. By Theorem 4.5, the pointwise stabilizer LLFL0L0F in H of {X,XF,e1,e1F} contains a non-scalar matrix D. The previous paragraph shows that D centralizes neither R nor S, and in fact that ZRZDZS.

Case (b): n = 5 and q = 4. Straightforward computations in Magma (with a runtime of about 1.5 h) show that |LLF| is even (for all FH), while |CH(R)|=625 (and hence |CH(S)|=625). It follows from the second last paragraph that ZRZDZS for each involution DLLF.

Case (c): n = 3. For each i{1,2,3}, define BiHZ as in Lemma 5.5. It is easy to check that [R,Bi]Z, and since Z, R and Bi are monomial matrices, Z,R,Bi<H. Thus ZRZBi. Similarly, ZSZBiF. Recall also from Lemma 5.5(ii)–(iii) that Bi,B1F stabilizes a one-dimensional subspace of V. Hence, as in the third last paragraph, if [Bi,B1F]1 for some i, then ZBiZB1F, and so ZRZBiZB1FZS. Otherwise, Lemma 5.5(iv) yields B1F=Bi for some i, and so ZRZBiZS.

 
Proof of Theorem 5.4

(i) Let x,yGZ(G). We will show that d(x,y)5. By Proposition 2.2, each of x and y is non-central in some maximal subgroup. Using Lemma 2.5, we may assume that each maximal subgroup K of G with xKZ(K) has odd order. By [35, Theorem 2], K=NH(S)/Z, where S is a Singer subgroup of H, that is, the intersection of H and a Singer subgroup of GL(n,q2). It follows from [30, p. 497 and p. 512] that NH(S)=S:C for some H-conjugate C of C(xn1).

Now, by Lemma 5.6(ii), ZCr for some involution rG, and so ZC,rLZ(L) for some maximal subgroup L of G of even order. As |C|=n is prime, each non-identity element of ZC lies in LZ(L). Since ZS, it follows that x=ZACi for some ASZ and some i{0,1,,n1}. Again using the fact that n is prime, it follows from Clifford’s Theorem that A acts irreducibly on Fq2n. Similar to Case (b) in the proof of Lemma 5.3 (but now working over Fq2), we see that [A,C]=Aq21. By Lemma 4.3, Aq41Z, and so [A,C]Z. Moreover, A,C,ZNH(S)<H. Therefore, ZAZC. It is now easy to see that ZACiZC for all i, and, in particular, xZC.

If yMZ(M) for some maximal subgroup M of G of even order, then since ZCr and |r|=2, Lemma 2.5 shows that d(ZC,y)4. Hence d(x,y)5. Otherwise, y lies in a G-conjugate of NH(S)/Z. By the previous paragraph, yZD for some H-conjugate D of C. Since xZC and d(ZC,ZD)3 by Lemma 5.6(iii), we again obtain d(x,y)5.

(ii) By [20, Remark 1.3] (see also [21, Ch. 4]), the intersection graph of G has diameter 5. Thus diam(Ξ(G))4 by Lemma 2.3. It remains to show that d(x,y)4 for all x,yG{1}. In what follows, except where stated otherwise, all information about G is determined using Magma, via computations involving conjugacy classes of elements and subgroups of G, with a runtime of about 2 min.

If |x|,|y|43, then each of x and y is adjacent in Ξ(G) to an involution, and Lemma 2.5 yields d(x,y)4. Assume from now on that |x|=43. Then x lies in no maximal subgroup of G of even order. As in the proof of (i), xt, where t is the image in G of some H-conjugate of C(x71). Since n > 3, it follows from Lemma 5.6(iii) that if |y|=43, then d(x,y)4.

Suppose finally that |y|43, and let T:=t. Then there exist maximal subgroups K and L of G such that yKZ(K), tLZ(L) and either

  • |KL|>|CG(t)|+|Z(K)|; or

  • Z(K)=1, CL(t)=T, TK=1 and |K||L|>|G|.

In case (I), there exists an element fKL that centralizes neither t nor K. Hence tf, and d(f,y)2 by Lemma 2.1, yielding d(t,y)3. In case (II), we observe that KL>1 and that tw for each non-identity wKL. As d(w,y)2 by Lemma 2.1, we again obtain d(t,y)3. We conclude in general that d(x,y)d(x,t)+d(t,y)4.

Using Magma, we observe that the non-commuting, non-generating graphs of PSU(3,3), PSU(3,4) and PSU(4,2) have diameters 2, 2 and 3, respectively.

5.3. The non-generating graph of a non-abelian finite simple group

Proof of Theorem 1.7

 

Let x,yG{1}, and let M1 and M2 be maximal subgroups of G containing x and y, respectively. If |M1| and |M2| are even, then a,b is dihedral for involutions aM1 and bM2. Hence (x,a,b,y) contains a path from x to y in Γ(G), and d(x,y)3. Thus we are done if every maximal subgroup of G has even order.

Assume now that |M1| is odd. Since Ξ(G) is a spanning subgraph of Γ(G), it follows from Theorems 2.6 and 1.1 that if G is not unitary of odd prime dimension, then diam(Γ(G)) is at most 5 if G=M, and at most 4 otherwise.

Suppose therefore that G=PSU(n,q), with n being an odd prime. As in the proof of Theorem 5.4, M1 is the image in G of the subgroup S:C of SU(n,q), with S being a Singer subgroup of SU(n,q) and C being a conjugate of the companion matrix C(xn1). By Lemma 5.6(ii), there exists an involution aG such that the image c of C in G is adjacent to a in Ξ(G), and hence in Γ(G).

Now, if M2 contains an involution b, then (x,c,a,b,y) contains a path from x to y in Γ(G) and d(x,y)4. Otherwise, by the previous paragraph, yc for some G-conjugate cʹ of c. Moreover, c and cʹ stabilize one-dimensional subspaces X and Xʹ, respectively, of Fq2n. Since q2 when n = 3, we deduce from Theorem 4.5 that GXGX contains an element t1. Hence (x,c,t,c,y) contains a path from x to y in Γ(G) and d(x,y)4. Therefore, diam(Γ(G))4.

It remains to determine lower bounds for diam(Γ(G)) when G{B,PSU(7,2),M}. Here, [20, Theorem 1.1 and Remark 1.3] and Proposition 3.10 show that the intersection graph of G is connected with diameter at least 5, and so Lemma 2.3 yields diam(Γ(G))4.

Acknowledgements

The author is grateful to Colva Roney-Dougal and Peter Cameron for helpful discussions regarding the original thesis on which this work is based and to Colva and anonymous referees for their useful comments on this paper.

Funding

This work was supported by the University of St Andrews (St Leonard’s International Doctoral Fees Scholarship and School of Mathematics and Statistics PhD Funding Scholarship) and EPSRC grant number EP/W522422/1.

References

1.

A.
Abdollahi
,
S.
Akbari
and
H. R.
Maimani
,
Non-commuting graph of a group
,
J. Algebra
298
 (
2006
),
468
492
.

2.

M.
Aschbacher
and
G. M.
Seitz
,
Involutions in Chevalley groups over fields of even order
,
Nagoya Math. J.
63
 (
1976
),
1
91
.

3.

W.
Bosma
,
J.
Cannon
and
C.
Playoust
,
The Magma algebra system. I. The user language
,
J. Symbolic Comput.
24
 (
1997
),
235
265
.

4.

T.
Breuer
,
The GAP Character Table Library, Version 1.3.9
(
2024
).

5.

T.
Breuer
,
R. M.
Guralnick
and
W. M.
Kantor
,
Probabilistic generation of finite simple groups. II
,
J. Algebra
320
 (
2008
),
443
494
.

6.

T.
Burness
,
R.
Guralnick
and
S.
Harper
,
The spread of a finite group
,
Ann. of Math. (2)
193
 (
2021
),
619
687
.

7.

T. C.
Burness
,
R. M.
Guralnick
and
J.
Saxl
,
On base sizes for algebraic groups
,
J. Eur. Math. Soc. (JEMS)
19
 (
2017
),
2269
2341
.

8.

T. C.
Burness
,
A.
Lucchini
and
D.
Nemmi
,
On the soluble graph of a finite group
,
J. Combin. Theory Ser. A
194
 (
2023
), 105708.

9.

P. J.
Cameron
,
Graphs defined on groups
,
Int. J. Group Theory
11
 (
2022
),
53
107
.

10.

P. J.
Cameron
,
S. D.
Freedman
and
C. M.
Roney-Dougal
,
The non-commuting, non-generating graph of a nilpotent group
,
Electron. J. Combin.
28
 (
2021
), 15,
Paper No. 1.16
.

11.

L.
Carlitz
,
Note on a quartic congruence
,
Amer. Math. Monthly
63
 (
1956
),
569
571
.

12.

J. H.
Conway
,
R. T.
Curtis
,
S. P.
Norton
,
R. A.
Parker
and
R. A.
Wilson
,
Atlas of Finite groups.
Oxford University Press
,
Eynsham
,
1985
.

13.

E.
Crestani
and
A.
Lucchini
,
The non-isolated vertices in the generating graph of a direct powers of simple groups
,
J. Algebraic Combin.
37
 (
2013
),
249
263
.

14.

B.
Csákány
and
G.
Pollák
,
The graph of subgroups of a finite group
,
Czechoslovak Math. J.
19
 (
1969
),
241
247
.

15.

W.
de Launey
and
D.
Flannery
,
Algebraic Design Theory
,
Amer. Math. Soc
,
Providence, RI
,
2011
.

16.

L. E.
Dickson
,
Linear groups: With an Exposition of the Galois Field theory.
Dover Publications Inc
,
New York, NY
,
1958
.

17.

H.
Dietrich
,
M.
Lee
,
A.
Pisani
and
T.
Popiel
,
Explicit construction of the maximal subgroups of the Monster
,
Preprint, 2024, arXiv:2411.12230
.

18.

H.
Dietrich
,
M.
Lee
and
T.
Popiel
.
The maximal subgroups of the Monster
,
Preprint
,
2023
,
arXiv:2304.14646
.

19.

S. D.
Freedman
.
Diameters of graphs related to groups and base sizes of primitive groups - GAP and Magma code (thesis data).
August
2022
, doi: .

20.

S. D.
Freedman
,
The intersection graph of a finite simple group has diameter at most 5
,
Arch. Math.
117
 (
2021
),
1
7
.

21.

S. D.
Freedman
,
Diameters of graphs related to groups and base sizes of primitive groups.
PhD thesis
,
University of St Andrews
,
May
2022
.

22.

S. D.
Freedman
,
The non-commuting, non-generating graph of a non-simple group
,
Algebr. Comb.
6
 (
2023
),
1395
1418
.

23.

T.
Fritzsche
,
The depth of subgroups of PSL(2,q)
,
J. Algebra
349
 (
2012
),
217
233
.

24.

The GAP Group, GAP – Groups, Algorithms and Programming, Version 4.12.2
(
2022
).

25.

G.
Glauberman
,
Central elements in core-free groups
,
J. Algebra
4
 (
1966
),
403
420
.

26.

D.
Gorenstein
,
R.
Lyons
and
R.
Solomon
,
The Classification of The Finite Simple groups. Number 3. Part I. Chapter A.
Amer. Math. Soc
,
Providence, RI
,
1998
.

27.

R. L.
Griess Jr.
,
Finite groups whose involutions lie in the center
,
Quart. J. Math. Oxford Ser. (2)
29
 (
1978
),
241
247
.

28.

G.
Heide
,
J.
Saxl
,
P. H.
Tiep
and
A. E.
Zalesski
,
Conjugacy action, induced representations and the Steinberg square for simple groups of Lie type
,
Proc. Lond. Math. Soc. (3)
106
 (
2013
),
908
930
.

29.

I. N.
Herstein
,
A remark on finite groups
,
Proc. Amer. Math. Soc.
9
 (
1958
),
255
257
.

30.

M. D.
Hestenes
,
Singer groups
,
Canadian J. Math.
22
 (
1970
),
492
513
.

31.

P. E.
Holmes
and
R. A.
Wilson
,
PSL2(59) is a subgroup of the Monster
,
J. London Math. Soc. (2)
69
 (
2004
),
141
152
.

32.

P. E.
Holmes
and
R. A.
Wilson
,
On subgroups of the Monster containing A5’s
,
J. Algebra
319
 (
2008
),
2653
2667
.

33.

B.
Huppert
,
Endliche Gruppen. I.
Springer-Verlag
,
Berlin-New York, NY
,
1967
.

34.

R.
Lidl
and
H.
Niederreiter
,
Finite fields.
CUP
,
Cambridge
,
1997
, 2nd Edn..

35.

M. W.
Liebeck
and
J.
Saxl
,
On point stabilizers in primitive permutation groups
,
Comm. Algebra
19
 (
1991
),
2777
2786
.

36.

M. W.
Liebeck
,
J.
Saxl
and
G. M.
Seitz
,
Subgroups of maximal rank in finite exceptional groups of Lie type
,
Proc. London Math. Soc. (3)
65
 (
1992
),
297
325
.

37.

M. W.
Liebeck
and
A.
Shalev
,
Simple groups, permutation groups and probability
,
J. Amer. Math. Soc.
12
 (
1999
),
497
520
.

38.

A.
Lucchini
and
D.
Nemmi
,
On the connectivity of the non-generating graph
,
Arch. Math.
118
 (
2022
),
563
576
.

39.

S.
Perlis
,
Theory of matrices
,
Dover Publications Inc
,
New York, NY
,
1991
.

40.

J. S.
Rose
,
On finite insoluble groups with nilpotent maximal subgroups
,
J. Algebra
48
 (
1977
),
182
196
.

41.

M.
Suzuki
,
On a class of doubly transitive groups
,
Ann. of Math. (2)
75
 (
1962
),
105
145
.

42.

R. A.
Wilson
,
The Finite Simple Groups
,
Springer-Verlag London, Ltd
,
London
,
2009
.

43.

R. A.
Wilson
. Maximal subgroups of sporadic groups,
Finite Simple Groups: Thirty Years of the Atlas and Beyond, Amer. Math.
Soc
.,
Providence, RI
,
2017
,
57
72
.

This is an Open Access article distributed under the terms of the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted reuse, distribution, and reproduction in any medium, provided the original work is properly cited.